P Ratio Affects Stability

Mar 28, 2019 - After polymer and pDNA cargo distribution was observed in vivo, P(EG-b-MAEMt) ... and protect genetic material offer a promising altern...
0 downloads 0 Views 5MB Size
Article Cite This: Biomacromolecules XXXX, XXX, XXX−XXX

pubs.acs.org/Biomac

Glycopolycation−DNA Polyplex Formulation N/P Ratio Affects Stability, Hemocompatibility, and in Vivo Biodistribution Haley R. Phillips,† Zachary P. Tolstyka,† Bryan C. Hall,‡ Joseph K. Hexum,† Perry B. Hackett,‡ and Theresa M. Reineke*,† †

Downloaded via UNIV AUTONOMA DE COAHUILA on March 28, 2019 at 12:05:55 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Center for Genome Engineering and Department of Chemistry, University of Minnesota, 207 Pleasant Street SE, Minneapolis, Minnesota 55455, United States ‡ Center for Genome Engineering and Department of Genetics, Cell Biology, and Development, University of Minnesota, Minneapolis, Minnesota 55455, United States S Supporting Information *

ABSTRACT: Genome editing therapies hold great promise for the cure of monogenic and other diseases; however, the application of nonviral gene delivery methods is limited by both a lack of fundamental knowledge of interactions of the gene-carrier in complex animals and biocompatibility. Herein, we characterize nonviral gene delivery vehicle formulations that are based on diblock polycations containing a hydrophilic and neutral glucose block chain extended with cationic secondary amines of three lengths, poly(methacrylamido glucopyranose-block-2-methylaminoethyl methacrylate) [P(MAG-b-MAEMt)-1, -2, -3]. These polymers were formulated with plasmid DNA to prepare polyelectrolyte complexes (polyplexes). In addition, two controls, P(EG-b-MAEMt) and P(MAEMt), were synthesized, formulated into polyplexes and the ex vivo hemocompatibility, or blood compatibility, and in vivo biodistribution of the formulations were compared to the glycopolymers. While both polymer structure and N/P (amine to phosphate) ratio were important factors affecting hemocompatibility, N/P ratio played a stronger role in determining polyplex biodistribution. P(EG-b-MAEMt) and P(MAEMt) lysed red blood cells at both high and low N/P formulations while P(MAG-b-MAEMt) did not significantly lyse cells at either formulation at short and medium polymer lengths. Conversely, P(MAG-b-MAEMt) did not affect coagulation at N/P = 5, but significantly delayed coagulation at N/P = 15. P(EG-b-MAEMt) and P(MAEMt) did not affect coagulation at either formulation. After polymer and pDNA cargo distribution was observed in vivo, P(EG-b-MAEMt) N/P = 5 and P(MAG-b-MAEMt) N/P = 5 both dissociated and deposited polymer in the liver, while pDNA cargo from P(MAG-b-MAEMt) N/P = 15 was found in the liver, lungs, and spleen. The contrast between P(MAG-bMAEMt) at N/P = 5 and 15 demonstrates that polyplex stability in the blood can be improved with N/P ratio and potentially aid polyplex biodistribution through simply varying the formulation ratios.



INTRODUCTION Genome editing holds great promise for the cure of monogenic diseases such as cystic fibrosis, Duchenne muscular dystrophy, hemophilia A and B, and epidermolysis bullosa.1−4 However, tissue nonspecificity limits the application of systemic genetic therapeutics, and many different strategies have been applied to both understand and improve material− tissue specificity at a fundamental level.2,5−8 Viral vectors provide efficient delivery of genetic cargo. However, packaging limitations, occasional safety issues such as immune response, and the high costs associated with mass production of clinicalgrade virus for treating large patient populations are major issues.9−11 Polymers with the ability to condense and protect genetic material offer a promising alternative to viral vectors. They are relatively easy to produce compared to viral vehicles, can safely package various plasmid sizes, and have shown significant uptake in a wide variety of human cell lines.2,6,8−10,12,13 Unfortunately, nonviral vehicles often © XXXX American Chemical Society

struggle to deliver therapeutic amounts of genetic material in vivo due to various systemic and intracellular transfection barriers including colloidal stability.6 Colloidal stability is an important delivery vehicle characteristic for preserving cargo circulation and bioavailability. Additionally, vehicle diameters must fall within a range of approximately 50−500 nm to avoid renal filtration and to achieve cellular endocytosis.14 Vehicles must avoid swelling or aggregation, which can lead to disassembly or precipitation. Swollen vehicles may expose their cargo to nucleases that can digest the genetic payload as well as target the complexes for kidney filtration.15 Aggregation enhances trapping of enlarged particles in lung capillary beds.2,15−17 Opsonization, an aggregation process with blood serum proteins, may produce Received: November 29, 2018 Revised: March 1, 2019

A

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

of targeting motifs, and accessible carbohydrate groups for cell surface receptor interaction.38 Glucose-based polymer delivery vehicles composed of poly(methacrylamidoglucopyranoseblock-2-methylaminoethyl methacrylate) (P(MAG-bMAEMt)) were previously designed that successfully displayed colloidal stability in media containing salt and serum and promoted gene expression in vitro.26,35,36 Herein, this study aims to characterize the stability, hemocompatibility, and in vivo biodistribution of P(MAG-bMAEMt)−DNA complexes (polyplexes) in human blood and in mice in an effort to further examine P(MAG-b-MAEMt) polyplex behavior in more complex, biologically relevant environments (Figure 1). We hypothesized that P(MAG-b-

particles that are actively removed by the body’s mononuclear phagocyte system (also called the reticuloendothelial system).2,18,19 Colloidal stability may also improve hemocompatibility. Hemocompatibility is defined in the International Standard (ISO) 10993-4, which is a document outlining how to test the blood compatibility of medical devices entitled: Biological evaluation of medical devices − Part 4: Selection of tests for interactions with blood. Per the standard, hemocompatible materials will avoid lysing red blood cells (hemolysis), activating the complement portion of the innate immune system, or altering blood coagulation properties (thrombogenicity).20 Strategic polymer design can aid refinement of delivery vehicles that avoid aggregation and contribute other desirable properties such as stable DNA-loading and protection, increased circulation time, low cytotoxicity, and high cellular uptake and consequent elevated gene expression.6,14 To condense and protect DNA, most polymer vehicles are cationic and bind anionic DNA via electrostatic interactions to form interpolyelectrolyte complexes (polyplexes).21,22 When polyplexes are formulated at an N/P ratio (the molar ratio of ionizable amines to ionizable phosphates in the polyplex system) greater than 1, there is a net positive charge, which increases association with negatively charged cell membranes and promotes endocytosis.2,23 However, this positive charge can also lead to nonspecific protein interactions and aggregation.2 In addition, higher N/P ratios often lead to increased cellular toxicity.24,25 Adding an uncharged hydrophilic block to a polycationic block forms a neutral and steric shell-like barrier around the cationic polyplex core to further protect the genetic material from degradation by nucleases, decrease cell exposure to positive charge and nonspecific uptake, and prevent polyplex aggregation and premature blood clearance.6,26 Polyplex formulations must be studied in detail so that formulations can be created that avoid aggregation and nonspecific tissue uptake, and promote hemocompatibility. Polyethylene glycol (PEG) is the most common neutral hydrophilic moiety used to increase polyplex biocompatibility, stability, and circulation time in vivo.27−30 PEG “brushes” as small as 2 kDa can block electrostatic interactions between negatively charged proteins and positively charged surfaces, and the attachment of PEG (i.e., PEGylation) to cationic polymers can prevent red blood cell hemolysis and decrease interaction with blood proteins such as fibrinogen and albumin.28,30,31 PEGylation of polymer−DNA complexes generally improves biocompatibility compared to the nonPEGylated counterparts; however, it has been shown to make liposomes susceptible to accelerated blood clearance (ABC) after multiple doses.32−34 The clearance is not likely due to autoantibody activity, but may be a result of either complement activation due to blood serum binding or IgM production against the PEGylated liposomes.32,34 As alternatives to PEG, our group has developed and studied biocompatible carbohydrate-based polymers that form colloidally stable, safe, and effective nonviral gene delivery vehicles.26,35−38 Multiple different sugars have been modified for use as monomers for Reversible Addition−Fragmentation Chain-Transfer (RAFT) polymerization and then used those monomers to design polymers able to deliver genetic material into several cell lines.26,36,39 By design, sugar moieties are often pendant to the polymer backbone, offering both hydroxyl groups for further chemical modification, such as the addition

Figure 1. Polymers are complexed with pDNA to form polyplexes at low and high N/P ratios. This study seeks to understand the role of polymer structure and the N/P ratio on polyplex biological properties. The results demonstrate that polymer structure more strongly influences hemocompatibility, while N/P ratio determines biodistribution.

MAEMt) polyplexes would improve hemocompatibility and biodistribution compared to the control polyplexes [P(EG-bMAEMt) and P(MAEMt)] due to the higher serum stability demonstrated previously and that the N/P ratio could significantly affect the results.35 This higher serum stability is likely a result of pendant P(MAG) molecules providing an improved steric barrier against protein aggregation compared to linear, smaller molecular weight PEG neutral blocks.26,36,37 It was found that P(MAG-b-MAEMt) polyplexes prevented the hemolysis of red blood cells for almost all formulations and B

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

(λ = 658 nm). Data were analyzed with Wyatt Technologies Astra VI software version 5.3.4.18 (Santa Barbara, CA). Fluorescent polymer tagging was quantified on a CUV 1 cm cuvette holder with a Mikropack DH-2000 Deuterium/Halogen TTC lamp from Ocean Optics Inc. (Dunedin, FL), and data were analyzed using Ocean Optics Inc. Basic Acquisition Software. DNA gels were visualized with a Spectroline BI-O-VISION Transilluminator (Westbury, NY) under the UV setting (312 nm) and were imaged using a Samsung Galaxy S3 camera phone. Polyplex size over time was determined using a Brookhaven BI-200SM DLS system. The system included a Brookhaven Mini L-30 HeNe laser (637 nm), a Brookhaven BI-NDO detector, a Brookhaven TurboCorr correlator, and a decalin bath. Hemoglobin absorbance was measured on a BioTek Synergy H1 hybrid plate reader (Winooski, VT) at 380, 415, and 450 nm. Morphology was determined by capturing differential interference contrast (DIC) images on an EVOS FL digital inverted microscope (Fisher Scientific, Pittsburgh, PA), while all animal and tissue imaging was conducted on an IVIS Spectrum with an excitation wavelength of 710 nm, an exposure time of 3 s, and an absorbance wavelength at 780 nm. Images were captured using Living Image software (PerkinElmer Inc., Waltham, MA). Animal tissue was bead milled with a BulletBlender from Laboratory Supply Network, Inc. (Atkinson, NH). Quantitative polymerase chain reaction (qPCR) experiments were performed using an Eppendorf Mastercycler (Hamburg, Germany). Fluorescent Tagging of Polymers. For in vivo studies, P(MAGb-MAEMt)-2 and P(EG-b-MAEMt) were labeled with a near-infrared fluorescent tag, Cyanine7-N-hydroxysuccinamide (Cy7-NHS), via amidation of the NHS ester to a pendant amine. The ratio of Cy7 to MAEMt functionality was 1:50 for both polymers; P(MAG-bMAEMt)-2: 0.00018 mmol Cy7, 0.0098 mmol MAEMt pendant moiety; P(EG-b-MAEMt): 0.00070 mmol Cy7, 0.035 mmol MAEMt pendant moiety. The reaction was carried out in a solution containing a 9:1 ratio of 0.1 M sodium acetate buffer to dimethylformamide (DMF). The mixture was vortexed and incubated for 4 h at 23 °C in the dark. The solutions containing the Cy7-labeled polymers were dialyzed exhaustively into DI water (3500 MWCO, pH adjusted to ∼5 with 1 M HCl) for 48 h and lyophilized, yielding a blue powder. The degree of labeling with Cy7 was measured by UV−vis (Ocean Optics, Inc.) absorbance at 750 nm using a solution of 0.2 mg/mL labeled polymer in water (Figure S5A, Supporting Information, P(MAG-b-MAEMt-2-Cy7): ε = 199 000, A750 = 0.166; P(EG-bMAEMt)-Cy7: ε = 199000, A750 = 0.355). Based on the absorbance measurements, the polymers were labeled at an efficiency of 1 Cy7 tag per 12.3 P(MAG-b-MAEMt)-2 polymer chains and 1 Cy7 to 9.7 P(EG-b-MAEMt) polymer chains. To verify that the fluorescence of the Cy7-tagged polymers was not quenched during polyplex formation, we compared the Cy7 fluorophore absorption of polyplexes to free Cy7-polymer (Figure S5B, Supporting Information). Polyplex Preparation, Binding, and Size. All polyplexes were formulated with pT2/CAL plasmid DNA (pDNA, 7537 bp) containing the firefly luciferase reporter gene.43,44 This plasmid includes a Sleeping Beauty transposon (pT2) and a CAGGS promoter that promotes relatively high gene expression in vertebrate cells.43,44 To formulate polyplexes, pDNA was diluted to 0.02 μg/μL in D5W. The overall volume varied based on the sample size needed for each experiment. The polymers were separately diluted with D5W to the same total volume as the pDNA. However, to achieve the appropriate N/P ratio, the concentration of polymer in solution varied according to the polymer molecular weight; the higher the N/ P ratio, the higher the concentration of cationic block in solution. P(MAG-b-MAEMt)-1, -2, -3, and P(EG-b-MAEMt) were each formulated at N/P = 5 and 15. Commercial controls, jetPEI and Glycofect, were formulated at N/P = 5 and 20, respectively. Following dilution, polymer and pDNA solutions were filtered separately through a 0.2 μm sterile syringe filter and then polymer solution was added to an equal volume of pDNA solution. Polyplexes were incubated at room temperature for 1 h before use.

did not visibly affect cell morphology. While most polyplex formulations were found to activate complement, vehicles with high amount (N/P = 15) of P(MAG-b-MAEMt) polymer delayed coagulation. Decreasing the N/P ratio to 5 restored normal coagulation. These in vitro coagulation results were the first indicator that tuning the polyplex formulations could affect blood compatibility and subsequent circulation. This finding was supported by the in vivo results that showed very different distribution patterns for P(MAG-b-MAEMt) polyplexes at N/P = 5 and 15. This striking difference in biodistribution between two polyplexes composed of the same polymer at different N/P ratios suggests not only that the N/P ratio has a stronger influence on biodistribution than polymer structure in this study, but also that biodistribution can potentially be tuned by varying polyplex formulation. The results described herein demonstrate in more detail the subtle differences in hemocompatibility and biodistribution between P(MAG-b-MAEMt), P(EG-b-MAEMt), and P(MAEMt) at relatively low and high N/P formulations.



MATERIALS AND METHODS

General. Materials. All chemicals or reagents were purchased from Sigma-Aldrich (St. Louis, MO) unless specified. Cyanine7 NHS ester (Cy7) was purchased from Lumiprobe (Hallandale Beach, FL). The control polymer jetPEI was purchased from Polyplus Transfection Inc. (Illkirch, France) and Glycofect was donated by Techulon (Blacksburg, VA). Dulbecco’s Modified Eagle’s Medium (DMEM), fetal bovine serum (FBS), phosphate buffered saline (PBS), and nuclease-free water were purchased from Gibco (Carlsbad, CA), and a sterile 5% dextrose by weight in water (D5W) solution was purchased through DME Supply Group (Augusta, GA). Agarose powder and 10× tris-acetate-ethylenediaminetetraacetic acid (TAE) gel supplies were obtained from Fisher BioReagents (Hampton, NH). Plasmids (pT2/CAL) were prepared as previously described.40−42 Blood, Biological Reagents, and Animals. Human whole blood with and without 3.2% sodium citrate anticoagulant was purchased from Memorial Blood Centers (St. Paul, MN). Protamine sulfate was obtained from MP Biomedicals (Santa Ana, CA). Partial thromboplastin time-Lupus anticoagulant reagent (PTT-LA), Neoplastin Plus, and 25 mM CaCl2 solution were purchased through Diagnostica Stago (Parsippany, NJ). The MBL oligomer kit (kit #029), the MicroVue C4d (kit #A008) and the MicroVue Bb Plus (kit #A027) kits were all purchased from Quidel (San Diego, CA), and the preactivated zymosan positive control was obtained through Complement Technology, Inc. (Tyler, TX). Ten-week-old C57BL/6 mice were obtained through the National Cancer Institute (NCI, Frederick, MD) and housed in pathogen-free conditions according to the Association for Assessment and Accreditation of Laboratory Animal Care (AAALAC) requirements and fed a normal diet. Mice were handled per the Institutional Animal Care and Use Committee (IACUC) using an Institutional Biosafety Committee (IBC) approved protocol (1202A09921). Lysis buffer was purchased from Qiagen (Hilden, Germany), and proteinase K was obtained through Roche (Basel, Switzerland). Instrumentation. 1H NMR spectra were obtained in D2O at 23 °C using a Bruker Avance III HD 500 MHz spectrometer from Cambridge Isotope Laboratories, Inc. (Andover, MA), and data were analyzed using Bruker Top Spin 3.1 software. Polymer dispersity was analyzed using size exclusion chromatography (SEC) on an Agilent 1260 High Performance Liquid Chromatography instrument (Santa Clara, CA) with 1.0 wt % acetic acid/0.1 M Na2SO4 eluent at a 0.4 mL/min flow rate. The system used Eprogen (Downers Grove, IL) size exclusion columns, [CATSEC1000 (7 μ, 50 × 4.6), CATSEC100 (5 μ, 250 × 4.6), CATSEC300 (5 μ, 250 × 4.6), and CATSEC1000 (7 μ, 250 × 4.6)], a Wyatt Dawn Heleos-II light scattering module (λ = 662 nm), and an Optilab T-rEX refractometer C

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules P(MAG-b-MAEMt)-1, -2, -3 polyplexes showed complete DNA complex formation at N/P = 5 according to a previous electrophoretic mobility shift assay.35 This gel-shift assay was performed with the P(EG-b-MAEMt) polyplexes as well as the complexes made with the Cy7-tagged polymers using P(MAG-b-MAEMt)-2 as a control. A 1.2% agarose gel was prepared by dissolving 1.2 g of agarose in 100 mL of 1× TAE buffer and the solution was heated for 30 min. The solution was then removed from the heat and 10 μL of ethidium bromide (EtBr) was mixed into the solution, which was cooled for 5 min. The gel was poured and allowed to solidify for 1 h. While the gel set, polyplex solutions in D5W were made for: pDNA only, jetPEI N/P = 5, Glycofect N/P = 20, P(EG-b-MAEMt) and P(MAG-b-MAEMt)-2 each at N/P = 5 and 15, P(EG-b-MAEMt)Cy7 at N/P = 5, and P(MAG-b-MAEMt)-2-Cy7 at N/P = 5 and 15. The polyplex solutions prepared at 0.02 μg/μL pDNA in D5W were allowed to complex for 1 h, then 45 μL of each sample was loaded into the gel. The gel was submerged in 1× TAE buffer and run at 125 mV for 1.5 h and then visualized with the transilluminator (Figure S6, Supporting Information). As an additional assessment of polyplex stability, 50 uL of each polyplex solution was mixed with 100 μL of 0.5 μg/mL EtBr in D5W. The fluorescence of each solution was measured by UV−vis (Ocean Optics, Inc.) at λ = 526 nm and compared to a pDNA only control that was also mixed with 0.5 μg/ mL EtBr in D5W. The results of this EtBr exclusion assay are shown in Figure S7, Supporting Information. For size and stability analysis, polyplexes were analyzed in D5W solution and in phenol red-free Dulbecco’s Modified Eagle’s Medium (DMEM) containing 10% fetal bovine serum (FBS) by volume (both at room temperature). The polyplexes samples analyzed were jetPEI N/P = 5; Glycofect N/P = 20; P(EG-b-MAEMt) at N/P = 5 and 15; and P(MAG-b-MAEMt)-1, -2, and -3 each at N/P = 5 and 15. Polyplex hydrodynamic diameters (HD) were measured by dynamic light scattering (DLS) over 2 h to determine if the polyplex sizes changed in this media.2,16,19 200 μL of each sample in D5W were analyzed for 10 min with a 637 nm laser at 20 mW using 90° backscatter detection, then 400 μL of DMEM/10% FBS was added to the polyplex solution. A measurement was taken 10 min after serum addition and recorded as t = 0. Measurements were repeated once each at 60 and 120 min. The correlation functions were processed into size distribution curves using REPES analysis, and further discussion on data fitting and analysis can be found in the Supporting Information (Figures S8−S20, Supporting Information). Peaks belonging to FBS protein (as compared to an FBS-only sample) were disregarded, and the polyplex peak in each sample is plotted in Figure 2. Hemolysis Assay and Morphology Analysis. Whole human blood in 3.2% sodium citrate anticoagulant was used to measure hemolysis. The whole blood was spun on an Eppendorf centrifuge (5810 R) at 2400g for 20 min, and the plasma supernatant was transferred to a 15 mL centrifuge tube and flash frozen for use in the coagulation assays. Once separated from the plasma, red blood cell concentrate (RCC) was either used fresh or stored at 4 °C for up to 30 days.43 The RCC was washed four times with ∼2 mL of PBS at room temperature, then diluted 1:9 in PBS (700 μL RC: 6300 μL PBS). Aliquots (900 μL) of 0.1× RCC were mixed with 100 μL of each polyplex sample in triplicate: jetPEI N/P = 5; Glycofect N/P = 20; P(EG-b-MAEMt) at N/P = 5 and 15; and P(MAG-b-MAEMt)-1, -2, and -3 each at N/P = 5 and 15. 100 μL of D5W solution was used as a negative control since D5W is isotonic with blood ex vivo, and one 5 μL aliquot of 0.1× RCC was added to 495 μL of sterile water as a 100% lysis positive control. All samples were incubated at 37 °C for 1 h. Following centrifugation at 3200g for 3 min at room temperature, 150 μL of the lysis positive control was pipetted into a clear 96-well plate (3 wells per sample). 100 μL of each sample was aliquoted into separate plate wells and then 50 μL of PBS was added to each sample. Absorbance was measured on a BioTek Synergy H1 hybrid plate reader at 380, 415, and 450 nm and the hemoglobin (Hb) concentration was calculated using the Harboe method (see eqs 1 and 2; DF = dilution factor).45,46 Equation 2 was used to calculate

Figure 2. Polyplex hydrodynamic diameters in D5W initially, then in DMEM plus 10% FBS over time. The x-axis specifies the polyplex formulation. Peak end points were recorded at t = 0, 1, and 2 h after addition of DMEM plus 10% FBS to the samples. Each data point represents one 10 min measurement. Note: The P(MAEMt) N/P = 5 and N/P = 15 particles at t = 1 and 2 h settled out of solution (based on visual observation) and their diameter could not be accurately plotted. percent hemolysis and the data were normalized to the 100% lysis control. Hb (g/L) = [(167.2 × A415) − (83.6 × A380) − (83.6 × A450 )] 1 1 × × (1) 1000 DF %Hemolysis =

Supernatant Hb (g/L) × 100 Total Hb (g/L)

(2)

To observe changes in red blood cell morphology with polyplex ̈ washed RCC was compared to RCC samples addition, naive incubated with polyplexes. All samples sat for 1 h at 37 °C, allowing cells to settle, and 1.25 μL of cells were drawn from the supernatant and further diluted into 3 μL PBS on a glass microscope slide, and topped with a coverslip. All images were collected using the differential interference contrast setting on an EVOS FL digital inverted microscope. Coagulation Assays. Platelet-poor plasma, hereon referred to as plasma, was separated from whole blood in 3.2% sodium citratecoated collection tubes following centrifugation at 2400g for 20 min at room temperature. Plasma was either used immediately or flash frozen in liquid nitrogen and stored at −80 °C for up to 30 days. Polyplex solutions (100 μL) in D5W were incubated with 900 μL of plasma for 1 h at 37 °C. D5W (100 μL) was used as a negative control for both the activated partial thromboplastin time (aPTT) and the prothrombin time (PT) assays. Data from either assay is a measure of the time it took for the sample plasma to coagulate and is reported in seconds. To observe coagulation in the aPTT assay, partial thromboplastin time-Lupus anticoagulant reagent (PTT-LA) reagent was prepared following the manufacturer’s instructions. 100 μL of PTT-LA was added to 100 μL of plasma sample mixture in a 35 mm Pyrex test tube in triplicate. Plasma incubated with D5W was the negative control, while the positive control was prepared by incubating plasma with protamine sulfate (3 mg/mL in PBS). While protamine sulfate is normally used to trigger coagulation, in this case we used it at high concentration to induce coagulation factor dilution, which delays coagulation.47 The tubes were warmed for 3−4 min in a 37 °C water bath. In a separate tube, 25 mM CaCl2 solution was warmed to 37 °C. Once all samples and reagents equilibrated to 37 °C, 100 μL of D

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

Figure 3. RBC lysis and morphology. (A) RBC lysis after 1 h of incubation with polyplexes. Hemolysis is proportional to supernatant hemoglobin absorbance and is presented as the percent of RBCs lysed (based on a 100% lysis sample by incubating RBCs in water). Data are the mean of three replicates ± standard deviation with the polymer abbreviation and N/P values on the x-axis. An asterisk indicates a statistical difference between the sample and the D5W negative control according to a Student’s t test with p < 0.05. (B) Representative morphology microscopy images of RBCs incubated with select polyplex samples. No distinctive differences in morphology can be seen. any additional reagent. To start coagulation, 200 μL of warm neoplastin was added to each sample tube, and PT was measured from time of reagent addition until visible clot formation. D5W in plasma was again the negative control, but the positive control was prepared by diluting plasma with 3× the normal amount of neoplastin reagent. This effectively delayed clotting by diluting the protein that forms clots, plasma fibrin, without completely preventing clot formation. Complement ELISA Assays. To screen the polyplex vehicles against complement activation, polyplexes were formulated in D5W as previously described and allowed to incubate 1 h at room temperature. Normal human serum was collected from donors through the Memorial Blood Center (St. Paul, MN) as whole blood in tubes lacking anticoagulant. The blood was allowed to clot at room temperature for 30 min, and the tubes were centrifuged at 1500g for 10 min. The serum was pipetted off the clot into 15 mL centrifuge tubes. Because it is projected that >10% of the general human population has a mannose-binding lectin (MBL) deficiency, we excluded serum samples lacking MBL prior to determining MBL activation for our samples.48 Serum (5 μL) from each donor was prescreened using an MBL oligomer ELISA kit purchased through

warm CaCl2 was added to a sample tube. The bottom of the tube was held in the 37 °C water bath to maintain consistent temperature, briefly lifted and tipped sideways for observation, then resubmerged into the bath. Time was recorded from the moment CaCl2 was added until a clot was seen in the tipped solution. This process was repeated twice more with each of the remaining warm samples for an n = 3. To test whether normal aPTT coagulation time could be restored, we ran a mixing assay where 50% more coagulation factor (i.e., plasma) was added to the system. In addition, we also tested whether the addition of more contact factor (PTT-LA reagent), would restore the coagulation time back to normal. The unclotted sample from the previously described coagulation assay was divided in half (to a volume of 150 μL). For the mixing assay, 1× more plasma (25 μL) was added to one of the test tubes. To the other tube was added 2× more PTT-LA reagent (50 μL). 50 μL CaCl2 was added to start coagulation, and clotting times were collected using the tipping method outlined above. PT was measured using the same manual tube tipping method, but with different reagents. Instead of CaCl2, neoplastin reagent was prepared per manufacturer instructions, then warmed to 37 °C. Plasma samples were prewarmed in the water bath for 1 min without E

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

Figure 4. Determination of polyplex effects on blood coagulation. (A) The effect of polyplex incubation as a function of N/P ratio (numbers on xaxis) on coagulation through the PT extrinsic pathway. (B) Coagulation time (s) of blood plasma samples via the aPTT intrinsic pathway after incubation with each polyplex sample for 1 h as a function of N/P ratio (numbers on x-axis). All data are presented as the mean number of seconds ± standard deviation with an asterisk denoting and variation from the negative control according to the Student’s t test with p < 0.05. Quidel, and it was found that all donor serum samples contained normal levels of MBL (data not shown). Each polyplex sample (10 μL) was then incubated in 90 μL human serum for 1 h at 37 °C while gently rocking. Two additional control samples were prepared by incubating 90 μL serum with either 10 μL D5W or 10 μL zymosan (10.0 mg/mL). The samples containing the serum and polyplexes were then screened for C4 and Bb complement protein content using the respective ELISA kits (Quidel) and following the manufacturer’s instructions. The ELISA absorbance values were converted to protein content (μg/mL) and plotted with the specific protein concentration on the y-axis. Biodistribution Studies. To observe biodistribution, P(MAG-bMAEMt)-2 and the control P(EG-b-MAEMt) polymer were fluorescently tagged with Cy7 as described above. These Cy7-tagged polymers formed Cy7-tagged polyplexes with pT2/CAL, which we used to fluorescently image whole mice and extracted organs. To avoid subjecting the mice to undue stress and hemotoxicity, we chose to test only one P(EG-b-MAEMt) formulation. While both the N/P = 5 and 15 formulations aggregated in serum and lysed red blood cells, the N/P = 15 formulation lysed a significantly higher percentage of RBC compared to the N/P = 5 formulation. Control mice were dosed with D5W only. For the glucose polymer, we observed that all

the polymers regardless of cationic block length behaved similarly with blood components, so we chose the midlength MAEMt block polymer, P(MAG-b-MAEMt)-2 to represent the glucose polymer series. We also knew from previous work that P(MAG-b-MAEMt)-2 gave the highest gene expression in HepG2 liver cells in vitro.35 To perform the biodistribution study, we formed polyplexes in D5W, allowed them to complex for 1 h at room temperature (as before), and injected 200 μL of the formulations into the tail vein of each mouse (pDNA concentration = 0.01 μg/μL). After euthanasia by CO2, all organ imaging was conducted on an IVIS Spectrum with an excitation wavelength of 710 nm, an exposure time of 3 s, and an absorbance wavelength of 780 nm. Measured fluorescence (Cy7) was expressed as radiant efficiency (p/cm2/s/sr)/(μW/cm2). Mice were euthanized approximately 30 min postinjection, and blood samples were drawn by cardiac puncture. Mice were then perfused with at least 20 mL of D5W. Mouse organs were collected, imaged within 30 min (within 1 h postinjection), and stored in liquid nitrogen. Fluorescent heat maps were generated using IVIS Living Image Software. All images were simultaneously loaded with the Living Image Browser and regions of interest (ROI), in this case individual organs, were isolated via the “Free Draw Method.” Each organ consisted of its own ROI where fluorescence is measured and F

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

Figure 5. Complement activation as assessed using ELISA screens for the production of pathway activation marker proteins. (A) C4 protein is a marker for CP and MBLP activation. Samples that significantly differ from D5W are marked with an asterisk. (B) Bb protein is an activation marker for the AP. Samples that significantly exceeded the D5W sample are marked with an asterisk. Each sample was run in duplicate and the error bars represent the standard deviation. Statistical markers were determined using one-way ANOVA (p < 0.05) followed by posthoc TukeyHSD tests.

Scheme 1. Characterization Summary for the Block Copolymers Synthesized for This Studya

a

(A) Structure of the P(MAEMtn) homopolymer. (B) Structure of P(MAGm-b-MAEMtn) polymers. (C) General structure of the P(EGm-bMAEMtn) polymer. (D) Polymer block degree of polymerization (DP) values for each polymer. The DP values for P(MAG-b-MAEMt)-1, -2, and -3 were determined using GPC while the P(EG-b-MAEMt) block lengths were determined using 1H NMR. Mn and Đ for all polymers were measured by GPC/light scattering. reported in radiant efficiency/ROI pixel area. Radiant efficiency is defined as the ratio of emission light over excitation light. Plotting the fluorescence using radiant efficiency over ROI pixel area accounts for

differences in both sample area and how the imager excitation light interacts with the irregular surface of the samples. Organ fluorescence was normalized to a background level of fluorescence selected from G

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules an area where there were no organs. This allowed organ fluorescence from different images to be compared and plotted on the same graph. To determine the in vivo location of the pDNA, we purified DNA from mouse organs and measured the amount of cargo plasmid delivered using quantitative polymerase chain reaction (qPCR). First, organs were homogenized using the Next Advance Bullet Blender according to an adaptation of the instrument manufacturer’s protocol for liver homogenization. Briefly, we placed the whole organ (or both organs, in the case of the lungs and kidneys) into an Eppendorf tube. Then we added 2−3 large stainless steel beads (1.6 mm diameter), about 20 smaller beads (0.9−2.0 mm blend), and 500 μL of lysis buffer (Qiagen) and ran the blender for 1 min. To extract the DNA, 20 μL (about 3 μg) of tissue homogenate from each organ sample was placed in a separate microcentrifuge tube with 665 μL lysis buffer and 15 μL Proteinase K (Roche) (20 μg/μL) and digested overnight at 55 °C on an elliptical rocker. DNA was isolated using a phenol chloroform protein extraction followed by ethanol DNA precipitation. Samples in a range of 200−1000 ng/μL of DNA were submitted to the University of Minnesota Genomics Core for qPCR analysis. The measurements were conducted using primers luc-F1 (sequence: TGAGTACTTCGAAATGTCCGTTC) and luc-R1 (sequence: GTATTCAGCCCATATCGTTTCAT). The fold increase in pDNA over background tissue readings was plotted. More experimental details about the biodistribution studies can be found in the Supporting Information. Statistical Analysis. All values reported are the mean ± standard deviation of measurements made in triplicate unless noted otherwise. All means for results in Figures 1−3 were compared to a specified control with the Student’s t test using JMP Pro Software (SAS Institute, Cary, NC) through the University of Minnesota Supercomputing Institute. All means in Figures 4−5 were compared via ANOVA followed by posthoc TukeyHSD test using R statistical software. Results were considered statistically different from the indicated control if p < 0.05.

sample and control polymers at N/P ratios of N/P = 5 and 15. The P(MAG-b-MAEMt)-1, -2, and -3 binding shift assay was previously published, but P(MAG-b-MAEMt)-2 at N/P = 5 and 15, jetPEI N/P = 5, Glycofect N/P = 20, and pT2/CAL only samples were also included as controls.35 The binding shift assay confirmed that the P(EG-b-MAEMt) polymer as well as the P(EG-b-MAEMt)-Cy7 and P(MAG-b-MAEMt)-2Cy7 polymers completely bound pT2/CAL plasmids as low as N/P = 5 (Figure S6). The polymers bound pT2/CAL plasmids at both N/P = 5 and 15 ratios and those formulations were used for further characterization.35 To further assess polyplex-binding capability, an ethidium bromide (EtBr) exclusion study was performed (Figure S7, Supporting Information). All polyplexes prevented EtBr from intercalating into the pDNA cargo as compared to a D5W only negative control, which suggests that all polymers were able to stably complex pDNA. For effective delivery to tissues in vivo, polyplexes should maintain a consistent size over time in the presence of serum proteins in the blood and physiological salt concentrations. Polyplex aggregation and opsonization can cause premature clearance from the bloodstream, immune response, and/or cause polyplexes to collect in the lung capillaries.2,6,15,18,19,27 Polyplex size was measured in the presence of DMEM/10% FBS solution over 2 h using DLS (see Figures S8−S20, Supporting Information, for the plot of each correlation function). Polyplexes prepared with jetPEI, Glycofect, and P(MAEMt) (all polymers lacking a protective neutral block) were found to aggregate upon incubation in the DMEM solution containing 10% FBS. Once protein was introduced to the system, P(MAEMt) and P(EG-b-MAEMt) polyplexes aggregated at both N/P = 5 and 15. In contrast, all polyplexes with a P(MAG) block maintained a relatively consistent hydrodynamic diameter of approximately 125 nm even upon addition of serum. All of the P(MAG-b-MAEMt) polyplex formulations were found to be colloidally stable, despite changes in N/P ratio and MAEMt block length, suggesting P(MAG) may be able to shield polyplexes from aggregation to a higher extent than PEG in vivo.26,49 DLS has several limitations that should be kept in mind when analyzing results. First, particle concentration must be kept low to avoid the signal from being scattered by more than one particle at a time. Second, large particles deflect a larger amount of light, which creates large amplitude signals that can drown out signals by smaller particles, even if there are more small particles than large particles. With the current DLS technique, it is difficult to differentiate particles that are similar in size, so polydisperse samples may appear to be monodisperse. Third, particles that settle out of solution cannot be analyzed by DLS. After 1−2 h of exposure to FBS proteins, P(MAEMt) N/P = 5 and 15 polyplexes aggregated to such a large size that they visibly settled out of solution, which is denoted using hash marks in Figure 2. Hemolysis and Morphology. Red blood cells (RBCs) compose the largest portion of whole blood and act as reliable indicators of toxicity in vivo.50 Accordingly, hemolysis and RBC morphology were examined to determine if polyplexes compromised RBC integrity. Polyplexes were formed in D5W as described above. Polyplex solution (100 μL) in D5W was incubated with 900 μL 0.1× RCC for 1 h at 37 °C, and the supernatant from each sample was measured for hemoglobin release, which is proportional to the amount of RBCs lysed.45,46 Polyplexes formed with all three P(MAG-b-



RESULTS AND DISCUSSION Polymer Synthesis and Characterization. The MAG and MAEMt monomers and the P(MAG-b-MAEMt) polymers were previously synthesized via RAFT polymerization according to our published method.35 The P(EG-b-MAEMt) control polymer was also synthesized via RAFT polymerization by using a PEG45-CTA and chain extending with MAEMt monomers using the same reaction conditions used to make the P(MAG-b-MAEMt) polymers. The polymers were characterized via GPC, 1H NMR, and SEC (Figures S1− S4, Supporting Information). Scheme 1 summarizes the characterization data obtained for the diblock copolymers investigated in this study. P(MAG-b-MAEMt) polymers consisted of a constant MAG block length and the length of the MAEMt cationic block was systematically increased from 30 to 76 repeat units. After Cy7-labeling P(EG-b-MAEMt) and P(MAG-bMAEMt)-2, the degree of labeling was assessed by measuring the fluorophore absorbance using UV−vis (Figure S5A). The degree of labeling was found to be approximately 1 Cy7 tag per 12 P(MAG-b-MAEMt)-2 polymer chains and approximately 1 Cy7 to 10 P(EG-b-MAEMt) polymers. To ensure the polymer complexation process did not quench the Cy7 signal, the fluorophore absorbance of free polymer and labeled polymer in polyplex form were compared using UV−vis (Figure S5B). There was no significant difference in absorbance between all polyplex formulations and the same concentration of free polymer in D5W, thus indicating that polymer formation does not quench the Cy7 signal. Polyplex Formation and Size Stability. An electrophoretic mobility shift assay was performed using all of the Cy7-tagged H

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

which plasma may clot, which were tested herein: the common, or intrinsic, pathway and the extrinsic pathway. As the pathway names suggest, the extrinsic pathway activates from external trauma that allows blood to escape the vascular system, while trauma internal to the vascular system triggers the intrinsic coagulation pathway through activation of platelets, collagen, and other biomarkers.59 An activated partial thromboplastin time (aPTT) test measures coagulation through the intrinsic pathway, and the prothrombin time (PT) assay tests the extrinsic pathway. Figure 4 displays clot formation monitored through PT and aPTT from plasma incubated with our polyplexes. The PT for all samples occurred within 12.5−14.5 s, a normal range by comparison to the D5W negative control. Hence the polyplexes did not affect extrinsic coagulation (Figure 4A). In the aPTT assay, all diblock polyplexes at N/P = 5 as well as jetPEI and Glycofect polyplexes presented normal coagulation times of about 40.0 s (Figure 4B). Interestingly, the polyplexes made with the P(MAG-bMAEMt) polymers at N/P = 15 significantly delayed coagulation time, while the block copolymer control, P(EGb-MAEMt), did not affect coagulation. The homopolymer, P(MAEMt), likewise did not alter aPTT coagulation. There are several known causes of aPTT delay: antibody production, unfractionated heparin treatment (UFH), coagulation factor deficiency, and contact factor deficiency.60 Indeed, antibodies are not a factor in these assays as all platelet-poor plasma (plasma) samples were prescreened before use to ensure they clotted in a normal time frame. For the same reason, plasma samples that may have been exposed to UFH were not used in these experiments. In this case, depletion of clotting factors in the plasma or depletion of contact factor in the thromboplastin reagent (Neoplastin Plus from Diagnostic Stago) are the two potential causes of coagulation delay. Others have shown that addition of PEG to surfaces that normally trigger coagulation on contact can prevent contact factor activation, although this mechanism cannot be fully supported by the current data.61 The experimental positive control, protamine sulfate, is a cationic polypeptide used to reverse heparininduced anticoagulation in the clinic at near 1:1 protamine:heparin ratios. However, when excess protamine is in the system, it has inherent anticoagulant activity by downregulating thrombin factor V activation through an unknown charge-mediated mechanism.62 This suggests that cationic moieties such as polyplexes at higher N/P ratios can delay coagulation through similar charge-mediated interactions with either plasma coagulation factor or the negatively charged contact factor included with the PTT-LA reagent (particulate silica, in this case). To test this hypothesis, a mixing study was run where key coagulation components were added back to plasma samples showing poor coagulation. Plasma samples incubated with polyplexes formulated with the block glycopolymers at N/P = 15 were split in half, either more plasma or more reagent (containing silica contact factor) was added to the test tube, and coagulation times were again recorded. If the polyplexes depleted coagulation factors or contact factors in the aPTT, the readdition of that component should return coagulation time to normal. Adding twice as much reagent to the samples decreased coagulation time for all samples, but only the addition of more plasma (i.e., more coagulation and contact factors) completely restored a normal aPTT (Figure S22, Supporting Information). This suggests that P(MAG-b-

MAEMt) N/P = 5 formulations, jetPEI, and Glycofect, did not lyse more cells than the negative control of RBCs incubated with D5W (Figure 3A). P(MAG-b-MAEMt)-1 at N/P = 15 also did not show lysis of red blood cells, but the polymer with the longest cationic block, P(MAG-b-MAEMt)-3, showed statistically higher red blood cell lysis at N/P = 15. This formulation has the highest ratio of cationic charge block to neutral glucose block, and it is known that particles with a net positive charge can disrupt cell membranes.51,52 However, it has also been shown that increasing the bulkiness of the monomers in the neutral block of the polymer or alternating the charged and neutral blocks within the polymer can decrease toxicity to cells.36,37,49 P(EG-b-MAEMt) and P(MAEMt) polyplexes at both N/P = 5 and 15 also demonstrated statistically significant RBC lysis compared to the D5W negative control. To observe possible changes in RBC morphology, visual observations were made for all samples diluted into PBS and imaged under a digital microscope (Figure 3B and Figure S21, Supporting Information). No significant morphological changes were noticed; all cells had a normal, disc-like shape with a dark central dimple. No lysed cells were visible using this method. These results combined with the DLS study demonstrate that P(MAG-b-MAEMt) polymers are colloidally stable in a physiological media model containing serum and in general do not cause destructive interactions with blood cells (with the exception of P(MAG-b-MAEMt)-3 at N/P = 15) over the experimental time course. The RBC lysis measured for P(MAEMt), P(EG-b-MAEMt), and P(MAG-b-MAEMt)-3 (N/P = 15) was likely caused by loss of cell membrane integrity after contact with the charged polyplexes, followed by an influx of solution. Notably, neither jetPEI nor Glycofect visually altered RBC morphology or caused lysis, nor did any of the P(MAG-b-MAEMt) polymers at low N/P ratios. Commercially available jetPEI is normally thought to be membrane lytic due to its high density of positive charge. However, to explain this observation, Godbey et al.53 showed that while free, linear PEI polymer can permeabilize membranes, the lytic effect decreases when the polymer is complexed with DNA.53 In a related study, Lambert and co-workers transfected human neural cells with PEI and confirmed that neural cell electrophysiological function was not affected, and suggesting that the plasma membrane was intact.54 While the PEI dose was toxic to cells, this did not coincide with increased membrane permeability. The lack of observed hemolysis by linear PEI further supports the conclusion that it did not lyse cell membranes when complexed to DNA at low concentrations. Coagulation. Plasma coagulation is an important blood function that can be altered even by polymers generally considered biocompatible (e.g., PEG). Heparin, for instance, is a common polyanion used to prevent coagulation and the formation of blood clots.55 Other polymers that are more structurally similar to our glycopolymer diblocks such as poly(amidoamine) (PAMAM) have been found to induce plasma clotting, while nanoparticles with a cationic galactose shell have been found to greatly delay coagulation.56−58 Furthermore, glycoproteins naturally activate coagulation through two different enzymatic pathways: the partial thromboplastin and the prothrombin pathways.59 With our polyplexes, plasma coagulation was tested following 1 h polyplex incubation to examine whether the delivery vehicles altered clot formation time. There are two pathways through I

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

Figure 6. Summary of biodistribution data. (A) Representative fluorescent mouse organ images for each sample exhumed at 1 h postinjection: (1) kidneys, (2) spleen, (3) liver, (4) lungs, (5) brain, (6) heart, (7) blood. (B) Organ fluorescence (excluding blood) in radiance efficiency was plotted by organ and polyplex sample. Each data point is the average of n = 5 samples (D5W samples have an n = 3), and error bars represent the standard deviation. An asterisk denotes samples that significantly differ from the D5W control according to two-factor ANOVA (p < 0.05). (C) DNA was extracted from the harvested organs (excluding blood) and analyzed by qPCR. Each data point is the average of n = 5 replicates (D5W samples have an n = 3). The amount of plasmid delivered to each organ was normalized to the amount of genomic background level to normalize for organ mass. Error bars represent the standard deviation. An asterisk denotes samples that significantly differ from the D5W control according to two-factor ANOVA (p < 0.05) followed by posthoc TukeyHSD tests.

MAEMT) N/P = 15 polyplexes have enough excess polymer to electrostatically bind negatively charged coagulation and contact factors in an ex vivo system. It is unclear whether these results could be directly extrapolated to an in vivo system having a responsive abundance of coagulation and contact

factors available, and caution should be used when drawing conclusions about the thrombogenic effects of polyplexes in vivo.61 Complement Activation. The complement system is a form of innate immunity in which foreign material can activate J

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

of the polyplexes, although it is currently unclear why this same mechanism would not decrease the complement activation of the P(MAEMt) polyplexes. Bb complement activation could also be caused by MAEMt, although additional testing with PEG and P(MAG) homopolymer controls would be needed to test this hypothesis. These data reveal that the P(MAG-b-MAEMt) block polymers have a similar profile to P(EG-b-MAEMt) polymers. Biodistribution. To promote in vivo delivery, polyplexes must circulate unobstructed through the vasculature to reach tissues of interest.2,16,19 During circulation, they must also avoid nonspecific protein aggregation, opsonization, disassembly, or immune system activation in order to increase the chance of target tissue uptake.2,14−17,71 One general strategy for avoiding nonspecific uptake and prolonging circulation is to incorporate hydrophilic block copolymers into delivery vehicles.72−74 The previously described experiments that suggest a lack of P(MAG-b-MAEMt) interaction with blood components for select formulations support our hypothesis that those formulations may remain intact during circulation. Because the polymer and genetic cargo are not covalently linked, the biodistribution of both the polymer and pDNA were measured individually. Formulations with P(MAG-bMAEMt)-2 at both N/P = 5 and 15 were selected for further study in vivo as the midlength MAEMt block gave the highest gene expression in previous cell studies.35 In addition, those formulations were found to be colloidally stable in salt/serum, hemocompatible, and display a similar complement activation profile as P(EG-b-MAEMt) control (used as the negative control at N/P = 5 as it caused less hemolysis than the same polyplexes at N/P = 15). Both P(MAG-b-MAEMt)-2 and P(EG-b-MAEMt) polymers were tagged with a Cy7 nearinfrared fluorophore to allow deep tissue imaging in organs, and were then complexed with pDNA to allow polyplex formation in D5W. Following tail vein injection of each polyplex formulations, mice were imaged for Cy7 fluorescence, then euthanized after 30 min. The polymer vehicle distribution to six major organs was assessed via fluorescence and the pDNA cargo distribution to the same organs was quantified (via tissue extraction and qPCR analysis of plasmid amount in each tissue). In addition to blood, the heart, brain, liver, lungs, kidneys, and spleen of each animal were collected. Tissues were imaged ex vivo (Figure 6A and S23, Supporting Information), and radiant efficiency was quantified for each tissue (Figure 6B). Background autofluorescence in the negative control (D5W only) mice was measured at approximately 5.0 × 108 for all organs (blood was not plotted due to low fluorescence in all samples). As shown in Figure 6B, P(EG-b-MAEMt) and P(MAG-b-MAEMt) N/P = 5 polymers were found in the liver. This suggests the polyplexes were not colloidally stable during circulation, and that the polymers were sequestered mostly in the liver during the experimental time period. Although the liver is the most active organ in the mononuclear phagocytosis system, the lack of signal in the macrophage-rich spleen suggests a potential separate mechanism of capture not solely related to phagocytosis by macrophages.66,75 It is possible that the P(MAG-b-MAEMt) N/P = 5 and P(EG-bMAEMt) polyplexes disassociated in the liver and were unable to circulate further, as has been demonstrated with PEGylated polyplexes previously.76 In fact, increasing the amount of P(MAG-b-MAEMt)-2 polymer in the formulation to an N/P ratio of 15 decreases polymer signal in the liver to background

complement proteins found in the blood and initiate an enzymatic cascade to remove that material. Complement activation can lead to material rejection, inflammation, antibody mediated injuries, and even cardiovascular distress due to hypersensitivity reactions.63,64 Therefore, it is crucial to test whether biomaterials activate complement pathways as a safety measure. To screen for complement activation, the polyplexes were incubated with human serum and then the amount of activated protein was quantified using an enzymelinked immunosorbent assay (ELISA) screen. The three different pathways through which complement activation can occur are the classical pathway (CP), the mannan-binding lectin pathway (MBLP), and the alternative pathway (AP).65−67 The CP and the MBLP partially overlap, and their activation correlates to the amount of complement component 4d (C 4d) generated, shown in Figure 5A. AP activation correlates to the amount of complement component Bb (Bb) shown in Figure 5B. All polyplexes were screened in both assays along with a D5W negative control and a positive control, zymosan, a polysaccharide derived from yeast known to activate all three complement pathways.68,69 As shown above, all block copolymers and homopolymers activated complement through either the CP or the MBLP (Figure 5A). Neither of the polymer controls (JetPEI or Glycofect) at the doses examined generated C4d protein, and therefore did not significantly activate the complement system via these two pathways in vitro. The CP is typically activated through contact with immune complexes such as immunoglobulins or with lipopolysaccharides.65,66 The MBLP can be activated by mannose polysaccharides.65 The results show that polyplexes formed with both the PEGylated and homopolymer controls along with the glucose-based polymers can activate the innate immune system, despite the fact that these polymer systems are not immunoglobulins or polysaccharides. The polymeric form of glucose could potentially mimic polysaccharides, thereby potentially contributing to immune recognition. Complement activation could also be caused by MAEMt through an unknown mechanism. Further work isolating the activation mechanism for these formulations could aid the development of polyplex vehicles that avoid CP and MBLP activation. The same polyplex formulations that generated C4d also generated significant levels of Bb, although the Bb levels did not approach that of the positive control, zymosan. This suggests that there is a spectrum of complement activation and that AP activation could be mitigated by cellular and humoral inhibitors in vivo.65 AP activation occurs through nonspecific covalent binding of complement protein to nucleophilic groups such as pathogen-associated molecular patterns (PAMPs) normally found on microorganisms.66 Hydroxyl groups on the surface of biomaterials amplify AP activity, which may be the mechanism by which the glycoplexes activate this pathway.70 Toda et al.67 demonstrated that amino groups can activate AP indirectly through proteins adsorbed on the material surface, although to a significantly lesser extent than activation by hydroxyl groups.67 This could be a possible activation mechanism for polyplexes that do not contain free −OH groups available for covalent binding to complement such as P(EG-b-MAEMt). Additionally, that same work found that positively charged surfaces adsorb other serum proteins that sterically prevent complement activation. This could be the mechanism by which jetPEI and Glycofect avoid activating the complement system as the charge is exposed on the surface K

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

encapsulation, circulation, and stability of polyplex formulations in vivo and may be a facile tunable handle to aid in vivo administration and biodistribution.

levels. It is striking that simply increasing the amount of polymer in the formulation 2-fold decreased the amount of polymer detected in the liver by an order of magnitude, indicating that the higher N/P formulation is more stable during circulation. Previous work has shown that about 10− 25% of polymer in a polyplex solution is free polymer; that is, the polymers are not complexed with the pDNA.77 This free polymer would likely show different biodistribution than complexed polymers and could explain why there is such a strong polymer signal in the liver, but low pDNA amounts in that same organ. However, the percentage of free polymer was about the same for N/P = 5 solutions as for N/P = 20 solutions made of the same polymer.77 If free polymer from a P(MAG-b-MAEMt)-2 N/P = 5 solution created a high signal in the liver, free polymer from an N/P = 15 solution should create an even higher signal in the liver. This further supports the hypotheses that N/P = 5 polyplexes dissociate in or before reaching the liver and that increasing the N/P to 15 changes polyplex biodistribution. Examining every possible polyplex fate is outside the scope of this study, but future studies to further understand the causes and mechanisms of P(MAG-bMAEMt) polyplex distribution and their effect on genetic cargo expression may prove useful. The amount of genetic material delivered to each organ was measured by qPCR (Figure 6C). The relative distribution of pDNA with the P(EG-b-MAEMt) and P(MAG-b-MAEMt)-2 N/P = 5 samples do not correlate with the polymer distribution shown in Figure 6B. Figure 6C shows that no significant levels of pDNA were detected in any of the organs for the P(EG-b-MAEMt) and P(MAG-b-MAEMt)-2 N/P = 5 formulations. This supports our hypothesis that these polyplexes disassociate in the liver and/or kidneys, leaving the pDNA to be cleared potentially through the waste (no bladder, urine, or fecal samples were collected in this study).76,78 The distribution of pDNA with P(MAG-bMAEMt)-2 N/P = 15 shows uptake in the liver, spleen, and lungs. The presence of DNA cargo in the liver and spleen suggests that macrophages may play a role in capturing the pDNA with the N/P = 15 polyplexes.66,75,76,78 Despite hemocharacterization and DLS data showing P(MAG-bMAEMt)-2 N/P = 15 to be stable against protein aggregation and relatively benign in human blood (both desired functions for promoting stealth circulation in vivo), this polyplex formulation appeared to still accumulate in the lungs after 30 min (both polymer carrier and pDNA), which could suggest aggregation of polyplexes during blood circulation that is structure dependent. Together, the data in Figure 6 suggest that polyplexes made at N/P = 5 are unstable before reaching cells that could endocytose them, which means they are unstable either during circulation or shortly after reaching the liver tissues. While P(MAG-b-MAEMt)-2 N/P = 15 appears to be more stable against decomplexation in the blood and potentially a better polyplex formulation for promoting circulation, highly charged polyplexes are still captured by phagocytic organs. Polyplex distribution to the lung may also indicate aggregation in the blood, which cannot be modeled with in vitro tests (vide infra). Indeed, these data indicate that while in vitro screens for colloidal stability and hemocompatibility are useful, they are not accurate models to predict both encapsulation and colloidal stability of polyplexes in vivo. Furthermore, higher N/P ratios do not necessarily need to be avoided for systemic administration; our results suggest that they promote



CONCLUSION In this work, the contribution of a hydrophilic glucose block to the hemocompatibility and biodistribution pattern of cationic polymer gene delivery vehicles ex vivo and in vivo was assessed. The glucose block copolycations, P(MAG-b-MAEMt)-1, -2, and -3, formed polyplex formulations with pDNA that were colloidally stable from aggregation in physiological salt and serum conditions as determined by DLS. P(MAEMt) and P(EG-b-MAEMt) were synthesized as controls, and unlike the glucose systems, aggregated (according to DLS measurements) under the same physiological salt and serum conditions. P(MAG-b-MAEMt)-1, -2, and -3 polyplex samples were exposed to human blood and did not lyse red blood cells (except for P(MAG-b-MAEMt)-3 N/P = 15) nor visibly affect cell morphology. In contrast, the P(MAEMt) and P(EG-bMAEMt) control polyplex formulations lysed cells even at the lower N/P = 5 ratio. All glucose block copolymers at N/P = 15 delayed aPTT coagulation by reversibly limiting coagulation factor availability, and both the control and glycopolymer formulations activated the complement system. While they did activate the complement system, P(MAG-bMAEMt)-2 N/P = 5 polyplexes were overall the most hemocompatible polyplex formulation according to hemolysis and coagulation results. Polymer structure and N/P ratios were found to affect polyplex−blood interactions, and both variables were further probed in vivo. Cy7-tagged polyplexes were administered to mice via the tail vein, and biodistribution of the polymer and pDNA distribution to different tissues was measured by fluorescent imaging and qPCR, respectively. Polymer from the P(EG-b-MAEMt) and P(MAG-b-MAEMt)2 formulations at N/P = 5 is found in the liver, with the pDNA cargo showing up at background levels in all organs. P(MAG-b-MAEMt) N/P = 15 polyplexes reversed this trend by showing background levels of polymer in all organs, with high levels of pDNA in the liver, lungs, and spleen. These data suggest that the lower N/P formulations disassembled either in the blood or the liver, leaving the pDNA to be excreted by the kidneys, while the higher N/P polyplexes were most likely captured by macrophages. The contrast between P(MAG-bMAEMt) at N/P = 5 and 15 shows that polyplex biodistribution can be tuned by varying the formulation ratios. In vitro studies identified polymer structure and polyplex formulation as determinant variables for blood interaction, but N/P ratio seems to play a stronger role than polymer structure in determining polyplex stability and distribution in vivo. Collectively, the hemocompatibility and biodistribution studies show that P(MAG-b-MAEMT)-2 polyplexes are more stable in serum solutions and cause less hemolysis than P(EG-bMAEMt) polyplexes but do affect coagulation when using higher polymer formulations. Polyplex formulation also affects in vivo biodistribution, with P(MAG-b-MAEMt)-2 N/P = 5 polyplexes behaving much differently than N/P = 15 vehicles composed of the same polymer and pDNA. It is possible that fine-tuning the glucose polymer structures and the polyplex formulations could create vehicles that avoid complement activation and allow circulation to organs beyond the liver, lungs, and spleen. Further work to measure circulation time, the effects of repeated dosing, and the unique distribution of these vehicles is underway. L

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules



(9) Walther, W.; Stein, U. Viral vectors for gene transfer: a review of their use in the treatment of human diseases. Drugs 2000, 60 (2), 249−71. (10) Thomas, C. E.; Ehrhardt, A.; Kay, M. A. Progress and problems with the use of viral vectors for gene therapy. Nat. Rev. Genet. 2003, 4 (5), 346−58. (11) Wu, Z.; Yang, H.; Colosi, P. Effect of Genome Size on AAV Vector Packaging. Mol. Ther. 2010, 18 (1), 80−86. (12) Uchida, E.; Mizuguchi, H.; Ishii-Watabe, A.; Hayakawa, T. Comparison of the efficiency and safety of non-viral vector-mediated gene transfer into a wide range of human cells. Biol. Pharm. Bull. 2002, 25 (7), 891−897. (13) Maurisse, R.; De Semir, D.; Emamekhoo, H.; Bedayat, B.; Abdolmohammadi, A.; Parsi, H.; Gruenert, D. C. Comparative transfection of DNA into primary and transformed mammalian cells from different lineages. BMC Biotechnol. 2010, 10, 9. (14) Rejman, J.; Oberle, V.; Zuhorn, I. S.; Hoekstra, D. Sizedependent internalization of particles via the pathways of clathrinand caveolae-mediated endocytosis. Biochem. J. 2004, 377 (Pt 1), 159−169. (15) Deshmukh, M.; Kutscher, H. L.; Gao, D.; Sunil, V. R.; Malaviya, R.; Vayas, K.; Stein, S.; Laskin, J. D.; Laskin, D. L.; Sinko, P. J. Biodistribution and renal clearance of biocompatible lung targeted poly(ethylene glycol) (PEG) nanogel aggregates. J. Controlled Release 2012, 164 (1), 65−73. (16) Dash, P. R.; Read, M. L.; Barrett, L. B.; Wolfert, M. A.; Seymour, L. W. Factors affecting blood clearance and in vivo distribution of polyelectrolyte complexes for gene delivery. Gene Ther. 1999, 6 (4), 643−650. (17) Alexis, F.; Pridgen, E.; Molnar, L. K.; Farokhzad, O. C. Factors affecting the clearance and biodistribution of polymeric nanoparticles. Mol. Pharmaceutics 2008, 5 (4), 505−515. (18) Shah, N. B.; Vercellotti, G. M.; White, J. G.; Fegan, A.; Wagner, C. R.; Bischof, J. C. Blood-nanoparticle interactions and in vivo biodistribution: impact of surface PEG and ligand properties. Mol. Pharmaceutics 2012, 9 (8), 2146−2155. (19) Ogris, M.; Brunner, S.; Schuller, S.; Kircheis, R.; Wagner, E. PEGylated DNA/transferrin-PEI complexes: reduced interaction with blood components, extended circulation in blood and potential for systemic gene delivery. Gene Ther. 1999, 6 (4), 595−605. (20) Buchannan, J. M.; Upman, P. J.; Wallin, R. F. (1998, November). A practical guide to ISO 10993-4: Hemocompatibility. MDDI. Retrieved from https://www.mddionline.com/practicalguide-iso-10993-4-hemocompatibility. (21) Katayose, S.; Kataoka, K. Remarkable increase in nuclease resistance of plasmid DNA through supramolecular assembly with poly(ethylene glycol)-poly(L-lysine) block copolymer. J. Pharm. Sci. 1998, 87 (2), 160−163. (22) Ruponen, M.; Ylä-Herttuala, S.; Urtti, A. Interactions of polymeric and liposomal gene delivery systems with extracellular glycosaminoglycans: physicochemical and transfection studies. Biochim. Biophys. Acta, Biomembr. 1999, 1415 (2), 331−341. (23) Yue, Z.-G.; Wei, W.; Lv, P.-P.; Yue, H.; Wang, L.-Y.; Su, Z.-G.; Ma, G.-H. Surface Charge Affects Cellular Uptake and Intracellular Trafficking of Chitosan-Based Nanoparticles. Biomacromolecules 2011, 12 (7), 2440−2446. (24) Grandinetti, G.; Ingle, N. P.; Reineke, T. M. Interaction of poly(ethylenimine)-DNA polyplexes with mitochondria: implications for a mechanism of cytotoxicity. Mol. Pharmaceutics 2011, 8 (5), 1709−1719. (25) Wang, F.; Bexiga, M. G.; Anguissola, S.; Boya, P.; Simpson, J. C.; Salvati, A.; Dawson, K. A. Time resolved study of cell death mechanisms induced by amine-modified polystyrene nanoparticles. Nanoscale 2013, 5 (22), 10868−10876. (26) Smith, A. E.; Sizovs, A.; Grandinetti, G.; Xue, L.; Reineke, T. M. Diblock glycopolymers promote colloidal stability of polyplexes and effective pDNA and siRNA delivery under physiological salt and serum conditions. Biomacromolecules 2011, 12 (8), 3015−3022.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.biomac.8b01704. Supplemental synthetic and characterization procedures; 1 H NMR spectra, UV−vis, SEC chromatograms, gel electrophoresis shift assays, dynamic light scattering; additional information about coagulation studies and biodistribution (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: 612-624-8042 (T.M.R.). ORCID

Theresa M. Reineke: 0000-0001-7020-3450 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge the NIH Director’s New Innovator Program (DP2OD006669) and the Camille and Henry Dreyfus Foundation for funding this work. UMN Bruker NMR spectrometer research reported in this publication was supported by the Office of the Director, National Institutes of Health of the National Institutes of Health under Award Number S10OD011952. We also acknowledge the financial support from NIH grants 1R01DK082516 and P01HD32652. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. The authors also thank the Fairview Hospital Special Coagulation Clinic and the University Genomics Core at the University of Minnesota for research support. We also acknowledge work done using the IVIS Spectrum in vivo imaging system at the University of Minnesota−University Imaging Centers, http://uic.umn.edu.



REFERENCES

(1) Hsu, C. Y. M.; Uludağ, H. Nucleic-acid based gene therapeutics: delivery challenges and modular design of nonviral gene carriers and expression cassettes to overcome intracellular barriers for sustained targeted expression. J. Drug Targeting 2012, 20 (4), 301−328. (2) Aied, A.; Greiser, U.; Pandit, A.; Wang, W. Polymer gene delivery: overcoming the obstacles. Drug Discovery Today 2013, 18 (21−22), 1090−1098. (3) Boudes, P. F. Gene therapy as a new treatment option for inherited monogenic diseases. Eur. J. Intern. Med. 2014, 25 (1), 31− 36. (4) Vanden Oever, M. J.; Tolar, J. Advances in understanding and treating dystrophic epidermolysis bullosa. F1000Prime Rep. 2014, 6, 35. (5) Ferrari, S.; Geddes, D. M.; Alton, E. Barriers to and new approaches for gene therapy and gene delivery in cystic fibrosis. Adv. Drug Delivery Rev. 2002, 54 (11), 1373−1393. (6) Pack, D. W.; Hoffman, A. S.; Pun, S.; Stayton, P. S. Design and development of polymers for gene delivery. Nat. Rev. Drug Discovery 2005, 4 (7), 581−93. (7) Jones, C. H.; Chen, C. K.; Ravikrishnan, A.; Rane, S.; Pfeifer, B. A. Overcoming nonviral gene delivery barriers: perspective and future. Mol. Pharmaceutics 2013, 10 (11), 4082−98. (8) Blanco, E.; Shen, H.; Ferrari, M. Principles of nanoparticle design for overcoming biological barriers to drug delivery. Nat. Biotechnol. 2015, 33 (9), 941−51. M

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules (27) Nelson, C. E.; Kintzing, J. R.; Hanna, A.; Shannon, J. M.; Gupta, M. K.; Duvall, C. L. Balancing cationic and hydrophobic content of PEGylated siRNA polyplexes enhances endosome escape, stability, blood circulation time, and bioactivity in vivo. ACS Nano 2013, 7 (10), 8870−8880. (28) Han, D. K.; Park, K. D.; Ryu, G. H.; Kim, U. Y.; Min, B. G.; Kim, Y. H. Plasma protein adsorption to sulfonated poly(ethylene oxide)-grafted polyurethane surface. J. Biomed. Mater. Res. 1996, 30 (1), 23−30. (29) Beletsi, A.; Panagi, Z.; Avgoustakis, K. Biodistribution properties of nanoparticles based on mixtures of PLGA with PLGA-PEG diblock copolymers. Int. J. Pharm. 2005, 298 (1), 233−241. (30) Singh, S.; Papareddy, P.; Mörgelin, M.; Schmidtchen, A.; Malmsten, M. Effects of PEGylation on membrane and lipopolysaccharide interactions of host defense peptides. Biomacromolecules 2014, 15 (4), 1337−1345. (31) Gon, S.; Fang, B.; Santore, M. M. Interaction of Cationic Proteins and Polypeptides with Biocompatible Cationically-Anchored PEG Brushes. Macromolecules 2011, 44 (20), 8161−8168. (32) Dams, E. T.; Laverman, P.; Oyen, W. J.; Storm, G.; Scherphof, G. L.; van Der Meer, J. W.; Corstens, F. H.; Boerman, O. C. Accelerated blood clearance and altered biodistribution of repeated injections of sterically stabilized liposomes. J. Pharmacol. Exp. Ther. 2000, 292 (3), 1071−1079. (33) Ishida, T.; Maeda, R.; Ichihara, M.; Irimura, K.; Kiwada, H. Accelerated clearance of PEGylated liposomes in rats after repeated injections. J. Controlled Release 2003, 88 (1), 35−42. (34) Wang, X.; Ishida, T.; Kiwada, H. Anti-PEG IgM elicited by injection of liposomes is involved in the enhanced blood clearance of a subsequent dose of PEGylated liposomes. J. Controlled Release 2007, 119 (2), 236−244. (35) Li, H.; Cortez, M. A.; Phillips, H. R.; Wu, Y.; Reineke, T. M. Poly(2-deoxy-2-methacrylamido glucopyranose)-b-Poly(methacrylate amine)s: Optimization of Diblock Glycopolycations for Nucleic Acid Delivery. ACS Macro Lett. 2013, 2 (3), 230−235. (36) Wu, Y.; Wang, M.; Sprouse, D.; Smith, A. E.; Reineke, T. M. Glucose-Containing Diblock Polycations Exhibit Molecular Weight, Charge, and Cell-Type Dependence for pDNA Delivery. Biomacromolecules 2014, 15 (5), 1716−1726. (37) Tolstyka, Z. P.; Phillips, H.; Cortez, M.; Wu, Y.; Ingle, N.; Bell, J. B.; Hackett, P. B.; Reineke, T. M. Trehalose-Based Block Copolycations Promote Polyplex Stabilization for Lyophilization and in Vivo pDNA Delivery. ACS Biomater. Sci. Eng. 2016, 2, 43−55. (38) Dhande, Y. K.; Wagh, B. S.; Hall, B. C.; Sprouse, D.; Hackett, P. B.; Reineke, T. M. N-Acetylgalactosamine Block-co-Polycations Form Stable Polyplexes with Plasmids and Promote Liver-Targeted Delivery. Biomacromolecules 2016, 17 (3), 830−840. (39) Sprouse, D.; Reineke, T. M. Investigating the effects of block versus statistical glycopolycations containing primary and tertiary amines for plasmid DNA delivery. Biomacromolecules 2014, 15 (7), 2616−2628. (40) Belur, L. R.; Podetz-Pedersen, K.; Frandsen, J.; McIvor, R. S. Lung-directed gene therapy in mice using the nonviral Sleeping Beauty transposon system. Nat. Protoc. 2007, 2 (12), 3146−3152. (41) Kali, G.; Georgiou, T. K.; Iván, B.; Patrickios, C. S. Anionic amphiphilic end-linked conetworks by the combination of quasiliving carbocationic and group transfer polymerizations. J. Polym. Sci., Part A: Polym. Chem. 2009, 47 (17), 4289−4301. (42) Xu, X.; Smith, A. E.; Kirkland, S. E.; McCormick, C. L. Aqueous RAFT Synthesis of pH-Responsive Triblock Copolymer mPEO−PAPMA−PDPAEMA and Formation of Shell Cross-Linked Micelles. Macromolecules 2008, 41 (22), 8429−8435. (43) Aronovich, E. L.; McIvor, R. S.; Hackett, P. B. The Sleeping Beauty transposon system: A non-viral vector for gene therapy. Hum. Mol. Genet. 2011, 20 (R1), R14−20. (44) Belur, L. R.; McIvor, R. S.; Wilber, A. Liver-directed gene therapy using the sleeping beauty transposon system. Methods Mol. Biol. 2008, 434, 267−276.

(45) Han, V.; Serrano, K.; Devine, D. V. A comparative study of common techniques used to measure haemolysis in stored red cell concentrates. Vox Sang. 2010, 98 (2), 116−123. (46) Noe, D. A.; Weedn, V.; Bell, W. R. Direct spectrophotometry of serum hemoglobin: an Allen correction compared with a threewavelength polychromatic analysis. Clin. Chem. 1984, 30 (5), 627− 630. (47) Darlington, D. N.; Delgado, A. V.; Kheirabadi, B. S.; Fedyk, C. G.; Scherer, M. R.; Pusateri, A. E.; Wade, C. E.; Cap, A. P.; Holcomb, J. B.; Dubick, M. A. Effect of hemodilution on coagulation and recombinant factor VIIa efficacy in human blood in vitro. J. Trauma 2011, 71 (5), 1152−1163. (48) Thiel, S.; Frederiksen, P. D.; Jensenius, J. C. Clinical manifestations of mannan-binding lectin deficiency. Mol. Immunol. 2006, 43 (1−2), 86−96. (49) Sizovs, A.; McLendon, P. M.; Srinivasachari, S.; Reineke, T. M. Carbohydrate polymers for nonviral nucleic acid delivery. Top. Curr. Chem. 2010, 296, 131−190. (50) Liu, Z. H.; Jiao, Y. P.; Wang, T.; Zhang, Y. M.; Xue, W. Interactions between solubilized polymer molecules and blood components. J. Controlled Release 2012, 160 (1), 14−24. (51) Grandinetti, G.; Smith, A. E.; Reineke, T. M. Membrane and Nuclear Permeabilization by Polymeric pDNA Vehicles: Efficient Method for Gene Delivery or Mechanism of Cytotoxicity? Mol. Pharmaceutics 2012, 9 (3), 523−538. (52) Fröhlich, E. The role of surface charge in cellular uptake and cytotoxicity of medical nanoparticles. Int. J. Nanomed. 2012, 7, 5577− 5591. (53) Godbey, W. T.; Wu, K. K.; Mikos, A. G. Poly(ethylenimine) and its role in gene delivery. J. Controlled Release 1999, 60 (2−3), 149−160. (54) Lambert, R. C.; Maulet, Y.; Dupont, J. L.; Mykita, S.; Craig, P.; Volsen, S.; Feltz, A. Polyethylenimine-mediated DNA transfection of peripheral and central neurons in primary culture: probing Ca2+ channel structure and function with antisense oligonucleotides. Mol. Cell. Neurosci. 1996, 7 (3), 239−246. (55) Hirsh, J.; Oates, J. A.; Wood, A. J. J. Heparin. N. Engl. J. Med. 1991, 324, 1565−1574. (56) Jones, C. F.; Campbell, R. A.; Franks, Z.; Gibson, C. C.; Thiagarajan, G.; Vieira-de-Abreu, A.; Sukavaneshvar, S.; Mohammad, S. F.; Li, D. Y.; Ghandehari, H.; Weyrich, A. S.; Brooks, B. D.; Grainger, D. W. Cationic PAMAM dendrimers disrupt key platelet functions. Mol. Pharmaceutics 2012, 9 (6), 1599−1611. (57) Ahmed, M.; Lai, B. F.; Kizhakkedathu, J. N.; Narain, R. Hyperbranched glycopolymers for blood biocompatibility. Bioconjugate Chem. 2012, 23 (5), 1050−1058. (58) Narain, R.; Wang, Y.; Ahmed, M.; Lai, B. F.; Kizhakkedathu, J. N. Blood Components Interactions to Ionic and Nonionic Glyconanogels. Biomacromolecules 2015, 16 (9), 2990−2997. (59) Palta, S.; Saroa, R.; Palta, A. Overview of the coagulation system. Indian J. Anaesth. 2014, 58 (5), 515−523. (60) Chng, W. J.; Sum, C.; Kuperan, P. Causes of isolated prolonged activated partial thromboplastin time in an acute care general hospital. Singapore Med. J. 2005, 46 (9), 450−456. (61) Hansson, K. M.; Tosatti, S.; Isaksson, J.; Wettero, J.; Textor, M.; Lindahl, T. L.; Tengvall, P. Whole blood coagulation on protein adsorption-resistant PEG and peptide functionalised PEG-coated titanium surfaces. Biomaterials 2005, 26 (8), 861−872. (62) Ni Ainle, F.; Preston, R. J. S.; Jenkins, P. V.; Nel, H. J.; Johnson, J. A.; Smith, O. P.; White, B.; Fallon, P. G.; O’Donnell, J. S. Protamine sulfate down-regulates thrombin generation by inhibiting factor V activation. Blood 2009, 114 (8), 1658−1665. (63) Szebeni, J.; Baranyi, L.; Savay, S.; Milosevits, J.; Bunger, R.; Laverman, P.; Metselaar, J. M.; Storm, G.; Chanan-Khan, A.; Liebes, L.; Muggia, F. M.; Cohen, R.; Barenholz, Y.; Alving, C. R. Role of complement activation in hypersensitivity reactions to doxil and hynic PEG liposomes: experimental and clinical studies. J. Liposome Res. 2002, 12 (1−2), 165−172. N

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules (64) Puttarajappa, C.; Shapiro, R.; Tan, H. P. Antibody-mediated rejection in kidney transplantation: A review. J. Transplant. 2012, 2012, 193724. (65) Palarasah, Y.; Nielsen, C.; Sprogoe, U.; Christensen, M. L.; Lillevang, S.; Madsen, H. O.; Bygum, A.; Koch, C.; Skjodt, K.; Skjoedt, M. O. Novel assays to assess the functional capacity of the classical, the alternative and the lectin pathways of the complement system. Clin. Exp. Immunol. 2011, 164 (3), 388−395. (66) Fujita, T. Evolution of the lectin−complement pathway and its role in innate immunity. Nat. Rev. Immunol. 2002, 2 (5), 346−353. (67) Toda, M.; Kitazawa, T.; Hirata, I.; Hirano, Y.; Iwata, H. Complement activation on surfaces carrying amino groups. Biomaterials 2008, 29 (4), 407−417. (68) Smith, M. C.; Pensky, J.; Naff, G. B. Inhibition of zymosaninduced alternative complement pathway activation by concanavalin A. Infect. Immun. 1982, 38 (3), 1279−1284. (69) Hamad, I.; Hunter, A. C.; Szebeni, J.; Moghimi, S. M. Poly(ethylene glycol)s generate complement activation products in human serum through increased alternative pathway turnover and a MASP-2-dependent process. Mol. Immunol. 2008, 46 (2), 225−232. (70) Arima, Y.; Kawagoe, M.; Toda, M.; Iwata, H. Complement activation by polymers carrying hydroxyl groups. ACS Appl. Mater. Interfaces 2009, 1 (10), 2400−2407. (71) Al-Dosari, M. S.; Gao, X. Nonviral gene delivery: principle, limitations, and recent progress. AAPS J. 2009, 11 (4), 671−681. (72) Akhtar, S.; Benter, I. Toxicogenomics of non-viral drug delivery systems for RNAi: Potential impact on siRNA-mediated gene silencing activity and specificity. Adv. Drug Delivery Rev. 2007, 59 (2−3), 164−182. (73) Guerrero-Cázares, H.; Tzeng, S. Y.; Young, N. P.; Abutaleb, A. O.; Quiñones-Hinojosa, A.; Green, J. J. Biodegradable polymeric nanoparticles show high efficacy and specificity at DNA delivery to human glioblastoma in vitro and in vivo. ACS Nano 2014, 8 (5), 5141−5153. (74) Magalhães, M.; Farinha, D.; Pedroso de Lima, M. C.; Faneca, H. Increased gene delivery efficiency and specificity of a lipid-based nanosystem incorporating a glycolipid. Int. J. Nanomed. 2014, 9, 4979−4989. (75) Zhang, Y.; Satterlee, A.; Huang, L. In vivo gene delivery by nonviral vectors: overcoming hurdles? Mol. Ther. 2012, 20 (7), 1298−1304. (76) Merkel, O. M.; Librizzi, D.; Pfestroff, A.; Schurrat, T.; Buyens, K.; Sanders, N. N.; De Smedt, S. C.; Béhé, M.; Kissel, T. Stability of siRNA polyplexes from poly(ethylenimine) and poly(ethylenimine)g-poly(ethylene glycol) under in vivo conditions: effects on pharmacokinetics and biodistribution measured by Fluorescence Fluctuation Spectroscopy and Single Photon Emission Computed Tomography (SPECT) imaging. J. Controlled Release 2009, 138 (2), 148−159. (77) Wang, X.; Kelkar, S. S.; Hudson, A. G.; Moore, R. B.; Reineke, T. M.; Madsen, L. A. Quantitation of complexed versus free polymers in interpolyelectrolyte polyplex formulations. ACS Macro Lett. 2013, 2, 1038−1041. (78) Zuckerman, J. E.; Choi, C. H. J.; Han, H.; Davis, M. E. Polycation-siRNA nanoparticles can disassemble at the kidney glomerular basement membrane. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (8), 3137−3142.

O

DOI: 10.1021/acs.biomac.8b01704 Biomacromolecules XXXX, XXX, XXX−XXX